ArticlePDF Available

Changes in Intracellular Calcium and Glutathione in Astrocytes as the Primary Mechanism of Amyloid Neurotoxicity

Authors:

Abstract and Figures

Although the accumulation of the neurotoxic peptide beta amyloid (betaA) in the CNS is a hallmark of Alzheimer's disease, the mechanism of betaA neurotoxicity remains controversial. In cultures of mixed neurons and astrocytes, we found that both the full-length peptide betaA (1-42) and the neurotoxic fragment (25-35) caused sporadic cytoplasmic calcium [intracellular calcium ([Ca2+]c)] signals in astrocytes that continued for hours, whereas adjacent neurons were completely unaffected. Nevertheless, after 24 hr, although astrocyte cell death was marginally increased, approximately 50% of the neurons had died. The [Ca2+]c signal was entirely dependent on Ca2+ influx and was blocked by zinc and by clioquinol, a heavy-metal chelator that is neuroprotective in models of Alzheimer's disease. Neuronal death was associated with Ca2+-dependent glutathione depletion in both astrocytes and neurons. Thus, astrocytes appear to be the primary target of betaA, whereas the neurotoxicity reflects the neuronal dependence on astrocytes for antioxidant support.
Content may be subject to copyright.
Changes in Intracellular Calcium and Glutathione in
Astrocytes as the Primary Mechanism of
Amyloid Neurotoxicity
Andrey Y. Abramov,
1
Laura Canevari,
2
and Michael R. Duchen
1
1
Mitochondrial Biology Group, Department of Physiology, University College London, London WC1E 6BT, United Kingdom, and
2
Miriam Marks Division
of Neurochemistry, Institute of Neurology, London WC1N 3BG, United Kingdom
Although the accumulation of the neurotoxic peptide
amyloid (
A) in the CNS is a hallmark of Alzheimer’s disease, the mechanism of
A neurotoxicity remains controversial. In cultures of mixed neurons and astrocytes, we found that both the full-length peptide
A
(1– 42) and the neurotoxic fragment (25–35) caused sporadic cytoplasmic calcium [intracellular calcium ([Ca
2
]
c
)] signals in astrocytes
that continued for hours, whereas adjacent neurons were completely unaffected. Nevertheless, after 24 hr, although astrocyte cell death
was marginally increased, 50% of the neurons had died. The [Ca
2
]
c
signal was entirely dependent on Ca
2
influx and was blocked by
zinc and by clioquinol, a heavy-metal chelator that is neuroprotective in models of Alzheimer’s disease. Neuronal death was associated
with Ca
2
-dependent glutathione depletion in both astrocytes and neurons. Thus, astrocytes appear to be the primary target of
A,
whereas the neurotoxicity reflects the neuronal dependence on astrocytes for antioxidant support.
Key words:
-amyloid; intracellular calcium; astrocyte; neuron; Alzheimer; glutathione
Introduction
Alzheimer’s disease (AD) is a neurodegenerative disorder char-
acterized by a progressive cognitive decline resulting from selec-
tive neuronal dysfunction, synaptic loss, and neuronal cell death.
It is accompanied by the deposition of the
-amyloid peptide
(
A), a polypeptide of 39 43 aa that is thought to play a major
role in the pathogenesis of the disorder (Small and McLean,
1999). Aggregated
A is neurotoxic. Although
A neurotoxicity
has been associated with oxidative stress and the reduction of
endogenous antioxidants (Behl et al., 1994; Casley et al., 2002a),
with mitochondrial damage (Casley et al., 2002b) and with the
destabilization of intracellular calcium ([Ca
2
]
c
) homeostasis,
both in neurons (Mattson et al., 1992) and in glial cells (Stix and
Reiser,1998), the mechanism of
A-induced neurotoxicity re-
mains uncertain.
Amyloid
peptides have been shown recently to form pores
in artificial membranes; it has been suggested that they may also
act as pore formers in intact neuronal membranes, in which they
appear to form Ca
2
-permeable channels (Arispe et al., 1993; Lin
et al., 2001).
A has also been shown to have effects on a variety of
types of ion-selective channels, including voltage-gated Ca
2
-
permeant channels (Blanchard et al., 1997; Ueda et al., 1997;
Rovira et al., 2002). More subtle changes in [Ca
2
]
c
signaling
have also been demonstrated after long-term exposure to
A
(Mattson and Chan, 2001), suggesting a disturbance of [Ca
2
]
c
homeostatic mechanisms that may reflect changes in cellular
metabolism.
Reports on the effects of
Aon[Ca
2
]
c
in astrocytes are con-
troversial. With exposure to
A, some authors have found that
astrocyte [Ca
2
]
c
increases (Stix and Reiser, 1998), whereas in
the hands of others, astrocyte [Ca
2
]
c
decreases (Meske et al.,
1998). The effect of
Aon[Ca
2
]
c
homeostasis in astrocytes is
potentially very important, considering the interplay between
neuronal and glial signals revealed recently (Haydon, 2001) and
the proposed glia-related pathomechanisms in AD (Harkany et
al., 2000; Schubert et al., 2001). Glial cells play a major supportive
role toward neurons, which includes supplying metabolic sub-
strates and the precursors of the antioxidant glutathione (GSH)
(Dringen, 2000) and removing excitatory amino acids such as gluta-
mate from the extracellular space (Takahashi et al., 1997), processes
that play a critical role in neuroprotection. These roles may be un-
dermined by the
A-induced generation of reactive oxygen species
(ROS) and the inhibition of glutamate uptake (Markesbery,1997;
Harkany et al.,2000), resulting in neuronal damage as a consequence
of impaired astrocytic support function.
Materials and Methods
Cell culture. Mixed cultures of hippocampal neurons and glial cells were
prepared as described previously (Vergun et al., 2001), with modifica-
tions, from Sprague Dawley rat pups 2– 4 d postpartum [University Col-
lege London (UCL) breeding colony]. Hippocampi were removed into
ice-cold Gey’s salt solution (Invitrogen, Paisley, UK) with 20
g/ml of
gentamicin. The tissue was minced and trypsinized (0.1% for 15 min at
37°C), triturated, and plated on poly-D-lysine-coated coverslips and cul-
tured in Neurobasal medium (Invitrogen) supplemented with B
27
(In-
vitrogen) and 2 mML-glutamine. Cultures were maintained at 37°C in a
humidified atmosphere of 5% CO
2
and 95% air, fed twice a week, and
maintained for a minimum of 10 d before experimental use to ensure the
expression of glutamate and other receptors. Neurons were easily distin-
guishable from glia: they appeared phase-bright, had smooth, rounded so-
Received Feb. 12, 2003; revised April 4, 2003; accepted April 10, 2003.
This work was supported by the Wellcome Trust, the Royal Society, and the Miriam Marks Foundation. We thank
Drs. Frances Edwards and Anna de Simoni for providing hippocampal explant cultures and Profs. S. Bolsover and J. B.
Clark and Drs. R. Dumollard and J. Jacobson for their invaluable discussion and suggestions.
Correspondence should be addressed to Michael R. Duchen, Department of Physiology, University College Lon-
don, Gower Street, London WC1E 6BT, UK. E-mail: m.duchen@ucl.ac.uk.
Copyright © 2003 Society for Neuroscience 0270-6474/03/235088-08$15.00/0
5088 The Journal of Neuroscience, June 15, 2003 23(12):5088 –5095
mata and distinct processes, and lay just above the focal plane of the glial
layer. Cells were used at 1020din vitro (DIV) unless stated otherwise.
Isolated cortical astrocytes were prepared as described previously
(Boitier et al., 1999). Cerebra taken from adult Sprague Dawley rats (UCL
breeding colony) were chopped and triturated until homogeneous,
passed through a 297
m mesh, and trypsinized (50,000 U/ml of porcine
pancreas; Sigma, Gillingham, UK) with 336 U/ml of DNase 1 (bovine
pancreas, Sigma), and 1.033 U/ml of collagenase (Sigma) at 37°C for 15
min. After the addition of fetal bovine serum (10% of final volume) and
filtering through 140
Mmesh, the tissue was centrifuged through 0.4 M
sucrose (400 gm, 10 min), and the resulting pellet was transferred to
Minimal Essential Medium supplemented with 5% fetal bovine serum, 2
mMglutamine, and 1 mMmalate in tissue culture flasks precoated with
0.01% poly-D-lysine. The cells reached confluence at 1214 DIV; they
were harvested and reseeded onto 24 mm diameter glass coverslips (BDH
Chemicals, Poole, UK) precoated with 0.01% poly-D-lysine for fluores-
cence measurements, and used over 24d.
Peptides and treatments.
A2535,
A142, and
A3525 (Bachem, St.
Helens, UK) were dissolved at 1 mMin sterile ultrapure water (Milli-Q stan-
dard; Millipore, Watford, UK) and kept frozen until use. The peptides were
added under the microscope, except for GSH and neurotoxicity measure-
ments, where they were added 24 hr before the experiment.
A2535 was
used at concentrations of up to 50
Mto ensure that it was present in molar
excess compared with inhibitors and so to exclude any direct interaction.
Microscopy. Fluorescence measurements were obtained using a Nikon
(Tokyo, Japan) epifluorescence inverted microscope with a 20fluorite
objective. Excitation light from a Xenon arc lamp is selected using 10 nm
bandpass filters centered at 340, 360, and 380 nm housed in a computer-
controlled filter wheel (Cairn Research, Faversham, UK). Emitted light
passed through a long-pass filter to a cooled CCD camera (Orca ER;
Hamamatsu, Welwyn Garden City, UK). All imaging data were collected
at intervals of 1015 sec, digitized, and analyzed using Kinetic Imaging
(Wirral, UK) software. Cells were protected from phototoxicity by inter-
posing a shutter in the light path to limit exposure between the acquisi-
tion of successive images.
Confocal images were obtained using a Zeiss (Oberkochen, Germany)
510 confocal laser scanning microscope and a 40oil immersion objec-
tive. The 488 nm argon laser line was used to excite fluo-4 fluorescence,
which was measured using a bandpass filter from 505 to 550 nm. Illumi-
nation intensity was kept to a minimum (at 0.1% of laser output) to avoid
phototoxicity, and the pinhole was set to give an optical slice of 2
m.
[Ca
2
]
c
measurements. Cells were loaded for 30 min at room temper-
ature with 5
Mfura-2 AM (Molecular Probes, Eugene, OR) and 0.005%
Pluronic in a HEPES-buffered salt solution composed of (in mM): 156
NaCl, 3 KCl, 2 MgSO
4
, 1.25 KH
2
PO
4
, 2 CaCl
2
, 10 glucose, and 10 HEPES,
pH 7.35. Traces, obtained using the cooled CCD imaging system, are
presented as ratios of excitation at 340 and 380 nm, both with emission at
515 nm. For some measurements, [Ca
2
]
i
was calculated using the
equation (Grynkiewicz et al., 1985): [Ca
2
]
c
K(RR
min
)/(R
max
R),
where Ris the fluorescence ratio (340/380 nm) and Kis the effective
dissociation constant of fura-2. R
max
and R
min
were determined by the
application of 50
Mdigitonin followed by 1 mMMnCl
2
.
All data presented were obtained from at least five coverslips and two
to three different cell preparations.
For confocal imaging, cells were loaded with fluo-4 AM (5
M; Molec-
ular Probes) for 20 min, followed by washing. Data are presented nor-
malized with respect to the first image of the sequence.
GSH measurements. To measure GSH, cells were incubated with 50
M
monochlorobimane (MCB; Molecular Probes) in HEPES-buffered salt
solution at room temperature for 40 min, or until a steady state had been
reached before images were acquired for quantitation (Keelan et al.,
2001). The cells were then washed with HEPES-buffered salt solution,
and images of the fluorescence of the MCB-GSH adduct were acquired
using the cooled CCD imaging system as described using excitation at
380 nm and emission at 400 nm.
Toxicity experiments. For toxicity assays we loaded cells simultaneously
with 20
Mpropidium iodide (PI), which is excluded from viable cells
but exhibits a red fluorescence after a loss of membrane integrity, and 4.5
MHoechst 33342 (Molecular Probes), which gives a blue stain to chro-
matin, to count the total number of cells. Using phase-contrast optics, a
bright-field image allowed the identification of neurons, which look
quite different from the flatter glial component and also lie in a different
focal plane, above the glial layer. A total of 600 800 neurons or glial cells
were counted in 20 25 fields of each coverslip. Each experiment was
repeated five or more times using separate cultures.
Statistical analysis. Statistical analysis and exponential curve fitting
were performed using Origin 7 (Microcal Software Inc., Northampton,
MA) software. Results are expressed as means SEM.
Results
We have used digital imaging techniques to explore the effects of
Aon[Ca
2
]
c
signals, antioxidant status, and cell viability in
cultures of mixed hippocampal neurons and astrocytes in cul-
tures ranging from 10 to 45 DIV. Application of either the full-
length peptide (142; 100 nMto 10
M) or the 2535 aa fragment
(150
M)of
A had no effect on [Ca
2
]
c
signals in neurons,
which remained quiescent over periods of up to 6 hr (n456
cells) (Fig. 1Aa,C). After 3040 min of incubation with
A the
cells showed robust and reversible responses to the application of
Figure 1.
Amyloid raises [Ca
2
]
c
in astrocytes and not in neurons. A, Records of fura-2
fluorescence from neurons ( a) and astrocytes (b) in hippocampal cocultures after exposure to
A25–35 peptide (50
M). The neurons showed no change in signal over a period of 35 min.
Their identity was confirmed by their response to glutamate (100
M) at the end of this period.
The astrocytes ( b) showed complex [Ca
2
]
c
fluctuations starting after 5– 6 min of exposure
to
A25–35 (50
M). These could continue for many hours. Some sample traces are extracted
from this population and illustrated as Bi,Bii, and Biii. The images in Care taken from a time
series of confocal images of a hippocampal coculture loaded with fluo-4. The field includes four
neurons (n) surrounded by astrocytes. Once again, the astrocytes show complex transient and
localized [Ca
2
]
c
responses whereas the neurons show no change in signal at all. The traces in
Doriginate from confocal images of a fluo-4-loaded hippocampal explant culture. Once again,
neuronsthatshowed a robust responsetoglutamate application showed no responseto
A(a)
whereas astrocytes showed complex [Ca
2
]
c
fluctuations and only a small transient metabo-
tropic [Ca
2
]
c
response to glutamate ( b). An image of a responding cell is inset for each signal.
Abramov et al. Astrocytic Calcium and GSH Changes in Amyloid Neurotoxicity J. Neurosci., June 15, 2003 23(12):5088 –5095 5089
glutamate (100
M) (Fig. 1Aa) or to depolarization with 50 mM
KCl, confirming both their viability and their neuronal identity.
Moreover, the presence of
A2535 or
A142 did not change
the amplitude of the [Ca
2
]
c
response to glutamate or to 50 mM
KCl; the response to the latter was 726 89 nM(n59 cells) in
control and 759 97 nM(n96 cells) in
A-incubated neurons.
Remarkably, astrocytes in the same cultures showed dramatic
[Ca
2
]
c
signals after exposure to
A, often occurring in cells
surrounding quiescent neurons (Fig. 1Ab,B,C) Similar [Ca
2
]
c
signals were seen in response to both the 2535 peptide and the
full-length 142 peptide, but no responses were seen to the re-
verse peptide 3525 (n135 cells). All [Ca
2
]
c
signals started
after a delay of 515 min and showed three patterns of re-
sponse: (1) sporadic increases in [Ca
2
]
c
, seen as low-amplitude
(100200 nM) [Ca
2
]
c
oscillations or fluctuations; (2) larger
spikes, followed by sustained elevated [Ca
2
]
c
(12
M), and (3)
very large increases in [Ca
2
]
c
, usually followed by loss of cell
viability. After washing the cells with
A-free saline, the re-
sponses persisted for up to 6 hr (data not shown).
Because responses like these have not been described previ-
ously, we were concerned that the properties of the cells might be
dictated by our culture conditions. Therefore, we repeated the
experiments using other culture systems. Experiments using cul-
tures of cortical astrocytes prepared in the same way gave results
identical to those from the hippocampus (data not shown). We
also used hippocampal explant cultures, in which the properties
of the tissue in vivo are well retained. Confocal imaging of explant
cultures loaded with fluo-4 again showed that exposure to
A2535 (50
M) provoked a selective increase in activity in glial
cells in the culture (Fig. 1 Db). Neurons were identified as the only
cells in the culture to show a rise in [Ca
2
]
c
with 50 mMKCl and
by their larger and more sustained response to glutamate (100
M) (Fig. 1Da). The astrocytes showed transient and oscillatory
activity with
A, showed no response to 50 mMKCl, and their
response to glutamate was a small transient response that reflects
activation of metabotropic glutamate receptors (n5 cultures).
Because the responses in dissociated cultures and explants ap-
peared similar, the dissociated cultures were used for the remain-
der of the experiments described.
Astrocyte [Ca
2
]
c
responses to
A are dependent on
extracellular Ca
2
and independent of intracellular
Ca
2
stores
The [Ca
2
]
c
responses to
A were never observed in Ca
2
-free
saline (n231 cells) (Fig. 2A). However, if cells were exposed to
AinaCa
2
-free saline and were then washed with
A-free,
Ca
2
-containing buffer, [Ca
2
]
c
responses were then seen in the
astrocytes (n69 cells) (Fig. 2B, bottom) whereas neuronal
[Ca
2
]
c
did not change beyond the small increase associated with
the restoration of basal Ca
2
influx (n64 cells) (Fig. 2B, top).
These data show that (1) the changes in [Ca
2
]
c
in astrocytes are
initiated through Ca
2
influx from external sources, (2) the ini-
tiation of the action of
A does not require the presence of Ca
2
,
and (3) the effect persists despite removal of
A from the saline.
The oscillatory [Ca
2
]
c
signals appeared typical of IP
3
-
mediated [Ca
2
]
c
release from endoplasmic reticulum (ER) seen
in astrocytes in response to a range of agonists (Peuchen et al.,
1996) and to mechanical stimulation (Charles et al., 1991).
Therefore, we considered whether external Ca
2
acts as a trigger
to activate phospholipase C (PLC), which would generate IP
3
and
so mobilize ER Ca
2
. However, the experiments illustrated in
Figure 3 suggest that PLC- and IP
3
-mediated signaling do not
play a significant role in the
A-induced [Ca
2
]
c
signals. Thus,
U73122 (5
M), an inhibitor of phospholipase C (Fig. 3A) did not
significantly impair the [Ca
2
]
c
astrocyte responses to
A(n
89 cells). Similarly, 2-APB, (40
M) an inhibitor of IP
3
-
dependent Ca
2
release, failed to reduce
A-induced [Ca
2
]
c
signals in astrocytes (n35 cells) (Fig. 3B), whereas it completely
blocked the [Ca
2
]
c
increase induced by ATP (100
M; data not
shown), which acts at purinergic receptors (P
2U
) to promote IP
3
-
dependent ER Ca
2
release (Peuchen et al., 1996).
Although the expression and role of ryanodine receptors
(RyRs) in astrocytes is debatable (Matyash et al., 2002), we also
tested the effect of the RyR inhibitor dantrolene (10
m), which
again had no significant effect on the
A-induced [Ca
2
]
c
signal
(n46 cells) (Fig. 3C). The incubation of astrocytes with 0.11
Mthapsigargin (an inhibitor of ER Ca
2
pumps) completely
depleted Ca
2
from the ER, demonstrated by the absence of a
[Ca
2
]
c
response to ATP (100
M) (Fig. 3D). The addition of
A
again then induced a [Ca
2
]
c
response that was not significantly
different from the control responses (Fig. 3D). In this instance,
values of peak [Ca
2
]
c
after thapsigargin were 390 54 nM,
compared with control responses to
A with peak values of 456
57 nM(n301 cells; p0.05). We also noted that the resting
Ca
2
level, which was usually slightly elevated after exposure to
thapsigargin because of the activation of store-operated Ca
2
influx, was slightly depressed by
A, suggesting that, if anything,
Figure 2. [Ca
2
]
c
responses to
A are dependent on extracellular Ca
2
.A, In the absence
of external Ca
2
, cortical astrocytes showed no change in [Ca
2
]
c
after exposure to
A25–35
(50
M). B, In a coculture exposed to
A25–35 (50
M) in the absence of external Ca
2
,no
response was seen in either neurons or astrocytes.
A was then washed out and external Ca
2
added. Despite the removal of the
A, the addition of Ca
2
caused a large increase in [Ca
2
]
c
in the astrocytes but only a very small change in the neurons, reflecting the restoration of basal
calcium entry. Once again, the neuronal identity was confirmed by the response to 100
M
glutamate at the end of the experiment.
5090 J. Neurosci., June 15, 2003 23(12):5088 –5095 Abramov et al. Astrocytic Calcium and GSH Changes in Amyloid Neurotoxicity
A suppresses store-operated Ca
2
influx. Taken together, these
data strongly suggest that ER-stored Ca
2
does not play a signif-
icant role in
A-induced [Ca
2
]
c
signals.
Mitochondria may have a high Ca
2
content in astrocytes
(Boitier et al., 1999), and we cannot overlook the possibility of the
participation of mitochondria in
A-induced [Ca
2
]
c
signals,
especially because A
causes mitochondrial damage (Casley et
al., 2002b). After the depolarization of mitochondria with the
uncoupler carbonyl cyanide p-trifluoromethoxyphenylhydra-
zone (FCCP; 1
M),
A2535 or
A142 still increased [Ca
2
]
c
(n84 cells). In the presence of FCCP, the [Ca
2
]
c
signals were
significantly increased (789 36 vs 456 58 nM;p0.005). This
could result from a fall in ATP or through the loss of mitochon-
drial Ca
2
uptake. Nevertheless, FCCP did not reduce the
A-
induced Ca
2
signal, showing that mitochondria cannot repre-
sent a significant source of Ca
2
.
Thus, taken together, these data strongly suggest that
A
causes [Ca
2
]
c
signals in astrocytes, but not in neurons, by in-
ducing a pathway for Ca
2
influx across the plasma membrane.
A induces Ca
2
influx into astrocytes
Extracellular Mn
2
enters cells via Ca
2
-permeant channels and
quenches the fluorescence of intracellular fura-2. This is most
readily seen when the fura-2 is excited at 360 nm, the Ca
2
-
independent (isosbestic) point of the fura-2 excitation spectrum,
whereas the Ca
2
-dependent change of the 340/380 nm ratio is
not altered. In the presence of
A(n154 cells), the signal
excited at 360 nm was unaltered in the absence Mn
2
, showing
that this reliably reports a Ca
2
-independent signal (Fig. 4A).
However, in the presence of 40
MMn
2
, the 360 nm fura-2
signal showed stepwise and irreversible decreases in the signal
corresponding with each transient increase in [Ca
2
]
c
(two exam-
ples are shown in Fig. 4B)
.
This approach also allowed us to test
whether
A caused Ca
2
entry in neurons that was masked by Ca
2
buffering. However,
A had no effect on the 360 nm fura-2 signal in
neurons (n120 cells) in the presence of Mn
2
, or, as shown above,
on the fura-2 ratio, confirming the selectivity of the action of
Aon
astrocytes. These observations suggest that each [Ca
2
]
c
transient
reflects a pulse of Ca
2
influx into the astrocytes.
Additional confocal imaging experiments showed that the
A-induced [Ca
2
]
c
signal was initiated as a rapid focal increase
in [Ca
2
]
c
. These responses could sometimes be seen clearly
originating from a point source (Fig. 5A,B), followed by slower
diffusion into the cytosol (Fig. 5A). In the examples in Figure 5B,
the rise in [Ca
2
]
c
was restricted to a small part of the cell and
failed to extend through the cytoplasm. This again is consistent
with the activation of an influx pathway followed by Ca
2
buff-
ering rather than the mobilization of ER stores, in which the
amplitude and rate of rise of the signal are maintained by active
propagation (Boitier et al., 1999).
Routes for Ca
2
influx
According to some authors (Ueda et al., 1997; He et al., 2002)
A
may induce a [Ca
2
]
c
signal in neurons by increasing Ca
2
influx
through voltage-dependent calcium channels (VDCC). Because as-
trocytes in our cultures do not show a [Ca
2
]
c
response to 50 mM
KCl, it seems unlikely that they express VDCC. Nifedipine (1
M),
an inhibitor of L-type VDCCs, had no effect on the shape or ampli-
tude of the
A-induced [Ca
2
]
c
signals (142 or 2535) in either
cortical or hippocampal astrocytes (n56 cells), but it completely
blocked the [Ca
2
]
c
response to 150 mMKCl in hippocampal neu-
rons. The responses were also not significantly affected by inhibitors
of either ionotropic or metabotropic glutamate receptors, including
20
MCNQX (n98 cells), 10
M()-5-methyl-10,11-dihydro-
5H-dibenzo [a,d] cyclohepten-5,10-imine maleate (n69 cells) or
50
M(S)-()-amino-4-carboxy-methyl-phenylacetic acid (n
178 cells) (data not shown), suggesting that the responses do not
reflect glutamate release into the culture.
An additional Ca
2
influx pathway expressed by glial cells and
probably not in neurons is the pathway for capacitative influx
(capacitative Ca
2
entry, CCE). One possibility is that
A acts
through altering the opening probability of this pathway. There-
fore, we tested the action of lanthanum, which blocks CCE (Pizzo
et al., 2001). However, La
3
(1
M) had no effect on the re-
sponses (n67 cells) (data not shown).
It has been suggested that in some cell types,
A-induced
increases in the generation of ROS serve as a trigger, which then
raise [Ca
2
]
c
(Varadarajan et al., 2000). The incubation of corti-
cal and hippocampal astrocytes with the antioxidant trolox (750
M) and ascorbate (1 mM, 45 min preincubation; n67 cells) or
the superoxide scavenger 4-hydroxy-2,2,6,6-tetramethyl-
piperadine-1-oxyl (500
M) plus catalase (250 U/ml) (n69
cells), which we have shown previously to be effective scavengers
of ROS (Vergun et al., 2001), did not have any significant impact
on the [Ca
2
]
c
response of the cells to
A (data not shown),
suggesting that the production of ROS by
A is not responsible
for the [Ca
2
]
c
increases in astrocytes.
Figure 3. Intracellular Ca
2
stores do not make a significant contribution to the astrocyte
[Ca
2
]
c
response to
A. Manipulations that either block components of the IP
3
and ryanodine
signaling pathways or that empty ER stores do not significantly alter the astrocyte responses to
A.Theapplicationof 50
M
A25–35caused[Ca
2
]
c
transientsincorticalastrocytes despite
thepresenceof5
MU73122(aninhibitorof PLC; A), 40
M2-APB ( B),dantrolene(an inhibitor
of ryanodine receptors; C), and 1
Mthapsigargin (an inhibitor of ER Ca
2
pumps; D), at a
concentration that prevented the response to ATP (100
M).
Abramov et al. Astrocytic Calcium and GSH Changes in Amyloid Neurotoxicity J. Neurosci., June 15, 2003 23(12):5088 –5095 5091
Zinc and clioquinol abolish the
A-
induced [Ca
2
]
c
response in astrocytes
A peptides have been shown to form chan-
nels in artificial and biological membranes
(Arispe et al., 1993; Lin et al., 2001; Kawa-
hara et al., 1997). Such channels can be
blocked by Zn
2
(Arispe et al., 1996). We
found that the incubation of hippocampal
or cortical astrocytes (n253 cells) with up
to1m
MZnCl
2
completely prevented the ef-
fect of
Aon[Ca
2
]
c
(Fig. 6A). However,
the addition of Zn
2
had no effect on the
A-induced [Ca
2
]
c
signals once they had
already started (data not shown), suggesting
that Zn
2
is not acting simply to block
channels, but rather to prevent their
formation.
A peptide binds to metal ions with a
selectivity Cu
2
Fe
3
Zn
2
(At-
wood et al., 1998), all of which promote
aggregation. The addition of Cu
2
(1
M
to1mM) did not change the amplitude
(456 58 to 490 56 nM;n81 cells) or
shape of the [Ca
2
]
c
signals in either cor-
tical or hippocampal astrocytes (Fig. 6B),
suggesting that endogenous heavy-metal
ions present in the culture are sufficient to
promote the aggregation of
A.
Cu
2
can undergo redox cycling and
generate ROS, whereas Zn
2
is not redox-
active but competes with Cu
2
for binding,
and therefore inhibits the oxidant proper-
ties of metal-bound
A (Cuajungco et al.,
2000). The inhibition of [Ca
2
]
c
signals by
Zn
2
does not appear to be dependent on
these redox properties, because: (1) we see
identical effects with both the full-length
peptide and the 2535 fragment, which does
not have the metal-binding coordination
site, and (2) we see no effect of antioxidants
on the responses (above). Trace amounts of
metals may promote aggregation in an
-helix conformation, whereas a high con-
centration of Zn
2
(and to a lesser extent,
Cu
2
) promote
-sheet fibrillar aggrega-
tion, which is classically associated with
A
toxicity. The channels blocked by Zn
2
have an
-helical structure, whereas the
mechanism described here seems more
likely to involve
-sheet formation because
(1) Zn
2
prevents but does not reverse the
response and (2)
A2535 cannot form
-helical channels but can form
-sheets.
Four types of conductances have been ob-
served with
A142 in lipid bilayers (Kou-
rie et al., 2001).
Clioquinol, a chelator of Cu
2
,Zn
2
,
and Fe
2
, prevents aggregation and res-
olubilizes
A; it has also been shown to have a beneficial effect in
mouse models of AD (Cherny et al., 2001; Melov, 2002). The
preincubation of cells with 12
Mclioquinol for 30 min dramat-
ically prevented the effect of
Aon[Ca
2
]
c
of cortical astrocytes
(n207 cells) (Fig. 6C).
A25–35 and
A1– 42 deplete GSH in hippocampal neurons
and astrocytes
We then explored the consequences of
A exposure for GSH,
using fluorescence imaging of the indicator MCB to identify
changes in GSH in different cell types within the same culture
Figure 4. Mn
2
quench confirms that astrocyte [Ca
2
]
c
transients reflect transient Ca
2
influx. Fura-2-loaded hippocampal
astrocytes showed typical [Ca
2
]
c
fluctuations (black line) in response to 50
M
A25–35. A, In the absence of external Mn
2
,
the fura-2 response excited at 360 nm (gray line) showed no change during the [Ca
2
]
c
transients, confirming that this is close to
the isosbestic [Ca
2
]
c
-independent excitation wavelength for fura-2. Bi, Bii, With the addition of 40
MMn
2
each [Ca
2
]
c
transient was accompanied by a step quench of the 360 nm fura-2 signal, confirming that each transient reflects a pulsed influx of
divalent cations seen in response to
A25–35.
Figure 5. Confocal imaging reveals focal Ca
2
influx in response to
A. In a hippocampal coculture loaded with fluo-4,
confocalimagingduring the exposure to
Ashowsthat the change in[Ca
2
]
c
canoriginateas a focal changethatdiffuses through
the cell and may be restricted to the subplasmalemmal space. Aa, Time series of confocal images taken during a single [Ca
2
]
c
transient response in an astrocyte. Note that the response begins with a focal rise in [Ca
2
]
c
(arrowhead) followed by the slower
spread through the cell. This is illustrated further in Ab, which shows a plot of the signal with time at four different locations in the
cell (indicated color-coded on the inset image). The rapid rate of rise at the point of influx contrasts with the much slower increase
seendeepinthe cytosol of the cell.B,Series of images taken fromanotherastrocyteduring a response to
A25–35,againshowing
that the [Ca
2
]
c
signal may be restricted to the periphery of the cell and fail to propagate through the cell. The first image of the
sequenceshowsthe raw data, whereasthesubsequent images showtheratio of the imagesequencewith respect tothefirst image
of the sequence.
5092 J. Neurosci., June 15, 2003 23(12):5088 –5095 Abramov et al. Astrocytic Calcium and GSH Changes in Amyloid Neurotoxicity
(Keelan et al., 2001). In agreement with previous reports (Casley
et al., 2002a; Muller et al., 1997; White et al., 1999), we found that
A significantly decreased GSH in cortical astrocyte monocul-
tures (by 54.7 4.9%; n798 cells; p0.001) (Fig. 7C) and in
hippocampal astrocytes in coculture (by 44.32 4.9%; control
100%) (Fig. 7A)(n831 cells; p0.005) after a 24 hr
incubation.
In contrast to the lack of effect of
A on neuronal [Ca
2
]
c
,
A
also significantly ( p0.01) reduced GSH in hippocampal neu-
rons (33.5 6.3%; n345 cells) (Fig. 7B). Because astrocytes
supply neighboring neurons with GSH precursors (Sagara et al.,
1993) the GSH decline in the neurons is
likely to be a secondary consequence of
the decrease of astrocyte GSH.
The removal of external Ca
2
alone
had no effect on GSH levels in hippocam-
pal neurons (n420 cells) or astrocytes
(n905 cells) in control experiments
(Fig. 7A–C). However, the effect of
A142 and
A2535 on GSH was abol-
ished in the absence of Ca
2
in both hip-
pocampal or cortical astrocytes and in
hippocampal neurons (Fig. 7A–C). Thus,
given the dependence of
A-induced
[Ca
2
]
c
fluctuations on external Ca
2
,it
seems likely that the GSH changes are a
direct consequence of the changes in as-
trocyte [Ca
2
]
c
.
In the presence of 12
Mclioquinol, a
concentration that abolished
A-induced
[Ca
2
]
c
fluctuations in astrocytes,
A2535 (and
A142, data
not shown) no longer caused a significant fall of GSH in cortical
astrocytes (n619 cells) (Fig. 7C).
Effect of
A on cell viability
We then examined the effect of a 24 hr exposure of cultures to
A2535 on cell viability and found that, remarkably, 49.9
8.5% of neurons (Fig. 8A) but only 23.2 4.2% of astrocytes
(n9 experiments) (Fig. 8B) died during this period in hip-
pocampal cocultures. Preincubation with 1
Mclioquinol re-
duced cell death of hippocampal neurons and of cocultured hip-
pocampal astrocytes by 50% (21.2 6.8% dead cells, p0.05,
and 15.2 3.2%, respectively, n5 experiments) (Fig. 8).
The removal of Ca
2
from the medium also significantly ( p
0.001) protected the hippocampal neurons (cell death fell from
49.9 8.5 to 16.45 4.1%; n6 experiments) and in cocultured
hippocampal astrocytes (from 23.2 4.2 to 17.5%; n7 exper-
iments; p0.05). The presence or absence of Ca
2
in the me-
dium did not change the percentage of dead cells in untreated
cells or in cells treated with the reverse peptide 3525 (Fig. 8 A,B).
Discussion
We have found that
A induces calcium signals selectively in
astrocytes causing sporadic fluctuations of [Ca
2
]
c
, while having
no apparent effect at all on [Ca
2
]
c
in nearby neurons. The
[Ca
2
]
c
signals are dependent on calcium influx from the extra-
cellular space and are inhibited by Zn
2
or by the heavy-metal
chelator clioquinol. Our data are most readily consistent with a
model in which
A inserts into the plasma membrane, in which it
either forms channels or influences the properties of an existent
Ca
2
-permeant channel. Several features of the responses are
remarkable in this respect: most notably, the selectivity of the
response for the astrocytes and the oscillatory, transient nature of
the [Ca
2
]
c
signals. One might anticipate that insertion of Ca
2
-
permeant channels into a cell membrane would generate a mono-
tonic increase in [Ca
2
]
c
, and the appearance of the transient
fluctuations were surprising. However, the Mn
2
quench and
confocal imaging data strongly argue that the transients do in-
deed result from transient episodic Ca
2
influx and so presum-
ably reflect either the sporadic openings of a channel with a low
opening probability or the transient formation of channels that
then dissociate.
The other major original findings reported here are that
A
affects [Ca
2
]
c
signals in astrocytes but not in neurons and that
Figure 6. Responses to
A are blocked by Zn
2
and clioquinol but not by Cu
2
.A, Addition of zinc (1 mMZnCl
2
) suppressed
the astrocyte response to
A. B, Addition of CuCl
2
(100
M) had no apparent effect on the
A responses, but the heavy-metal
chelator clioquinol (2
M)(C) suppressed the responses completely.
Figure 7.
A causes Ca
2
-dependent depletion of GSH in both neurons and astrocytes.
MCBwasused to image astrocyteandneuronal GSH by digitalimaging.Hippocampal cocultures
(15–20 DIV) (A, B) and cortical astrocyte cultures ( C) were treated for 24 hr with
A25–35 or
A35–25 (50
Mfor both) and clioquinol (1
M) at 37°C in culture medium with (gray col-
umns) or without (black columns) calcium. Mean intensities of MCB–GSH adduct fluorescence
(arb.U) are presented.
A25–35 decreased GSH dramatically either in hippocampal astrocytes
in coculture with neurons ( A) or in cortical astrocytes in monoculture (C). The response was
dependent on extracellular calcium and was also suppressed by clioquinol ( C).
A35–25 had
no effect. Note that neuronal GSH was also significantly reduced ( B) and that the reduction was
also calcium dependent, although it represents a proportionately smaller response than that of
the astrocytes.
Abramov et al. Astrocytic Calcium and GSH Changes in Amyloid Neurotoxicity J. Neurosci., June 15, 2003 23(12):5088 –5095 5093
the ensuing neurotoxicity appears to be secondary to impaired
astrocytic function in the support of neuronal viability, although
we cannot exclude a direct toxic effect of the
A on neurons. In
most published studies of
A neurotoxicity, either neuronal cell
lines were used (Blanchard et al., 1997), or, in studies of neurons
in primary culture, the presence and contribution of glial cells
was not excluded (Mattson et al., 1992; Pike et al., 1993). The
selectivity of the effects on [Ca
2
]
c
for astrocytes is remarkable
and may reflect some difference in the plasma membrane lipid
composition of the two cell types, because even small differences
in the lipid environment affect
A binding to membranes and
pore formation (Curtain et al., 2003). The stability of
A aggre-
gates in membranes is very delicately balanced (McLaurin and
Chakrabarty, 1996). For example, increased membrane choles-
terol, such as is found in AD brain and aging, favors a
-sheet
over an
-helix conformation of
A (Curtain et al., 2003). To our
knowledge, very little information is available on the membrane
composition in different brain-cell types. Alternatively, the selec-
tivity may reflect the selective effects of
A on an existing chan-
nel, which is expressed in astrocytes but not in neurons, although
our pharmacological search has failed to reveal such a process.
In conclusion, our experimental conditions have allowed us to
uncover a novel toxic mechanism of
A, probably overlapping in
vivo with other known effects. This involves
-aggregation of
A
and the selective insertion into the astrocyte plasma membrane,
initiating sporadic [Ca
2
]
c
signals, which can persist over long
periods. These signals, although not causing astrocyte cell death,
nevertheless promote GSH depletion in both cell populations;
they ultimately impair neuronal viability because GSH depletion
leaves the neurons vulnerable to damage by oxidative stress.
Thus, the resulting neurotoxicity reflects the central role of astro-
cytes in supporting neuronal function by supplying GSH precur-
sors and other metabolic intermediates and by removing excess
glutamate from the extracellular medium.
References
Arispe N, Rojas E, Pollard HB (1993) Alzheimer disease amyloid
protein
forms calcium channels in bilayer membranes: blockade by
tromethamine and aluminum. Proc Natl Acad Sci USA 90:567571.
Arispe N, Pollard HB, Rojas E (1996) Zn
2
interaction with Alzheimer
amyloid
protein calcium channels. Proc Natl Acad Sci USA
93:17101715.
Atwood CS, Moir RD, Huang X, Scarpa RC, Bacarra NM, Romano DM,
Hartshorn MA, Tanzi RE, Bush AI (1998) Dramatic aggregation of Alz-
heimer A
by Cu(II) is induced by conditions representing physiological
acidosis. J Biol Chem 273:1281712826.
Behl C, Davis JB, Lesley R, Schubert D (1994) Hydrogen peroxide mediates
amyloid
protein toxicity. Cell 77:817822.
Blanchard BJ, Konopka G, Russell M, Ingram VM (1997) Mechanism and
prevention of neurotoxicity caused by
-amyloid peptides: relation to
Alzheimers disease. Brain Res 776:4050.
Boitier E, Rea R, Duchen MR (1999) Mitochondria exert a negative feed-
back on the propagation of intracellular Ca
2
waves in rat cortical astro-
cytes. J Cell Biol 145:795808.
Casley C, Land J, Sharpe M, Clark J, Duchen M, Canevari L (2002a)
-Amyloid fragment 2535 causes mitochondrial dysfunction in primary
cortical neurons. Neurobiol Dis 10:258267.
Casley CS, Canevari L, Land JM, Clark JB, Sharpe MA (2002b)
-Amyloid
inhibits integrated mitochondrial respiration and key enzyme activities.
J Neurochem 80:91100.
Charles AC, Merrill JE, Dirksen ER, Sanderson MJ (1991) Intercellular sig-
naling in glial cells: calcium waves and oscillations in response to mechan-
ical stimulation and glutamate. Neuron 6:983992.
Cherny RA, Atwood CS, Xilinas ME, Gray DN, Jones WD, McLean CA,
Barnham KJ, Volitakis I, Fraser FW, Kim Y, Huang X, Goldstein LE, Moir
RD, Lim JT, Beyreuther K, Zheng H, Tanzi RE, Masters CL, Bush AI
(2001) Treatment with a copper-zinc chelator markedly and rapidly in-
hibits beta-amyloid accumulation in Alzheimers disease transgenic mice.
Neuron 30:665676.
Cuajungco MP, Goldstein LE, Nunomura A, Smith MA, Lim JT, Atwood CS,
Huang X, Farrag YW, Perry G, Bush AI (2000) Evidence that the
-amyloid plaques of Alzheimers disease represent the redox-silencing
and entombment of a
by zinc. J Biol Chem 275:1943919442.
Curtain CC, Ali FE, Smith DG, Bush AI, Masters CL, Barnham KJ (2003)
Metal ions, pH and cholesterol regulate the interactions of Alzheimers
disease amyloid-
peptide with membrane lipid. J Biol Chem
278:29772982.
Dringen R (2000) Metabolism and function of glutathione in brain. Prog
Neurobiol 62:649671.
Grynkiewicz G, Poenie M, Tsien RY (1985) A new generation of Ca
2
indi-
cators with greatly improved fluorescence properties. J Biol Chem
260:34403450.
Harkany T, Abraham I, Timmerman W, Laskay G, Toth B, Sasvari M, Konya
C, Sebens JB, Korf J, Nyakas C, Zarandi M, Soos K, Penke B, Luiten PG
(2000) beta-amyloid neurotoxicity is mediated by a glutamate-triggered
excitotoxic cascade in rat nucleus basalis. Eur J Neurosci 12:27352745.
Haydon PG (2001) Glia: listening and talking to the synapse. Nat Rev Neu-
rosci 2:185193.
He LM, Chen LY, Lou XL, Qu AL, Zhou Z, Xu T (2002) Evaluation of
-amyloid peptide 2535 on calcium homeostasis in cultured rat dorsal
root ganglion neurons. Brain Res 939:6575.
Kawahara M, Arispe N, Kuroda Y, Rojas E (1997) Alzheimers disease amy-
loid
-protein forms Zn
2
-sensitive, cation-selective channels across ex-
cised membrane patches from hypothalamic neurons. Biophys J
73:6775.
Keelan J, Allen NJ, Antcliffe D, Pal S, Duchen MR (2001) Quantitative im-
aging of glutathione in hippocampal neurons and glia in culture using
monochlorobimane. J Neurosci Res 66:873884.
Kourie JI, Henry CL, Farrelly P (2001) Diversity of amyloid
protein frag-
ment [140]-formed channels. Cell Mol Neurobiol 21:255284.
Lin H, Bhatia R, Lal R (2001) Amyloid
protein forms ion channels: impli-
cations for Alzheimers disease pathophysiology. FASEB J 15:24332444.
Figure 8.
A causes Ca
2
-dependent cell death in neurons and not in astrocytes. Effect of
A on viability of neurons and astrocytes. PI fluorescence was used to detect dead cells 24 hr
after the addition of 50
M
A and 1
Mclioquinol in the presence or absence of Ca
2
. Dead
cellswerecounted with respect to thetotalnumber of cells present,identifiedbystaining nuclei
with Hoechst 33342.
A caused a dramatic increase in cell death in neurons and only a modest
increase in astrocyte cell death in hippocampal cocultures. Cell death was calcium dependent
and cells were dramatically protected by 1
Mclioquinol.
5094 J. Neurosci., June 15, 2003 23(12):5088 –5095 Abramov et al. Astrocytic Calcium and GSH Changes in Amyloid Neurotoxicity
Markesbery WR (1997) Oxidative stress hypothesis in Alzheimers disease.
Free Rad Biol Med 23:134147.
Mattson MP, Chan SL (2001) Dysregulation of cellular calcium homeostasis
in Alzheimers disease: bad genes and bad habits. J Mol Neurosci
17:205224.
Mattson MP, Cheng B, Davis D, Bryant K, Lieberburg I, Rydel RE (1992)
-Amyloid peptides destabilize calcium homeostasis and render human
cortical neurons vulnerable to excitotoxicity. J Neurosci 12:376389.
Matyash M, Matyash V, Nolte C, Sorrentino V, Kettermann H (2002) Re-
quirement of functional ryanodine receptor type 3 for astrocyte migra-
tion. FASEB J 16:8486.
McLaurin J, Chakrabartty A (1996) Membrane disruption by Alzheimer
-amyloid peptides mediated through specific binding to either phos-
pholipids or gangliosides: implications for neurotoxicity. J Biol Chem
271:2648226489.
Meske V, Hamker U, Albert F, Ohm TG (1998) The effects of
/A4-amyloid
and its fragments on calcium homeostasis, glial fibrillary acidic protein
and S100
staining, morphology and survival of cultured hippocampal
astrocytes. Neuroscience 85:11511160.
Melov S (2002) . . . and C is for Clioquinol: the A
Cs of Alzheimers dis-
ease. Trends Neurosci 25:121123.
Muller WE, Romero FJ, Perovic S, Pergande G, Pialoglou P (1997) Protec-
tion of flupirtine on
-amyloid-induced apoptosis in neuronal cells in
vitro: prevention of amyloid-induced glutathione depletion. J Neuro-
chem 68:23712377.
Peuchen S, Clark JB, Duchen MR (1996) Mechanisms of intracellular cal-
cium regulation in adult astrocytes. Neuroscience 71:871883.
Pike CJ, Burdick D, Walencewicz AJ, Glabe CG, Cotman CW (1993) Neu-
rodegeneration induced by
-A peptides in vitro: the role of peptide
assembly state. J Neurosci 13:16761687.
Pizzo P, Burgo A, Pozzan T, Fasolato C (2001) Role of capacitative calcium
entry on glutamate-induced calcium influx in type-I rat cortical astro-
cytes. J Neurochem 79:98109.
Rovira C, Arbez N, Mariani J (2002) A
(2535) and A
(140) act on
different calcium channels in CA1 hippocampal neurons. Biochem Bio-
phys Res Commun 296:13171321.
Sagara J, Miura K, Bannai S (1993) Maintenance of neuronal glutathione by
glial cells. J Neurochem 61:16721676.
Schubert P, Ogata T, Marchini C, Ferroni S (2001) Glia-related patho-
mechanisms in Alzheimers disease: a therapeutic target? Mech Ageing
Dev 123:4757.
Small DH, McLean C (1999) Alzheimers disease and the amyloid-
pro-
tein: what is the role of amyloid? J Neurochem 73:443449.
Stix B, Reiser G (1998)
-Amyloid peptide 2535 regulates basal and
hormone-stimulated Ca
2
levels in cultured rat astrocytes. Neurosci Lett
243:121124.
Takahashi M, Billups B, Rossi D, Sarantis M, Hamann M, Attwell D (1997)
The role of glutamate transporters in glutamate homeostasis in the brain.
J Exp Biol 200:401409.
Ueda K, Shinohara S, Yagami T, Asakura K, Kawasaki K (1997) Amyloid
protein potentiates Ca
2
influx through L-type voltage-sensitive Ca
2
channels: a possible involvement of free radicals. J Neurochem
68:265271.
Varadarajan S, Yatin S, Aksenova M, Butterfield DA (2000) Alzheimers
amyloid
-peptide-associated free radical oxidative stress and neurotox-
icity. J Struct Biol 130:184208.
Vergun O, Sobolevsky AI, Yelshansky MV, Keelan J, Khodorov BI, Duchen
MR (2001) Exploration of the role of reactive oxygen species in gluta-
mate neurotoxicity in rat hippocampal neurones in culture. J Physiol
(Lond) 531:147163.
White AR, Bush AI, Beyreuther K, Masters CL, Cappai R (1999) Exacerba-
tion of copper toxicity in primary neuronal cultures depleted of cellular
glutathione. J Neurochem 72:20922098.
Abramov et al. Astrocytic Calcium and GSH Changes in Amyloid Neurotoxicity J. Neurosci., June 15, 2003 23(12):5088 –5095 5095
... Specifically, Abramov et al. demonstrated that calcium accumulation occurs not as a monotonic increase but in short waves [50,51]. In these studies, which were performed on individual cells in Figure 13. ...
... After the content of the lysosome is leaked completely, the recovery will be swift due to the presence of homeostatic mechanisms controlling intracellular pH and calcium concentrations. Correspondingly, for ions that are endocytosed and reach lysosomes but are not controlled by homeostatic mechanisms, such as manganese, the intracellular concentration will increase synchronously to the rising front of calcium waves (as shown in Figure 14, A) but would not return to baseline [51]. ...
... This question was asked before. In the study's design, Abramov et al. hypothesized that intracellular ion disturbances arise from the appearance of non-selective membrane ion channels formed by beta-amyloid [51]. The formation of such channels was indisputably described by multiple laboratories starting in 1993 [53][54][55][56][57][58][59][60][61][62][63]. ...
Preprint
Full-text available
The manuscript presents the comprehensive integrative theory of the etiology and pathogenesis of Alzheimer’s disease - the amyloid degradation toxicity hypothesis - and describes the logic that underlies it.The analysis of amyloid biomarkers and stable-isotope label kinetics (SILK) studies suggest that AD diagnosis is associated with higher cellular uptake of beta-amyloid. Uptake of beta-amyloid by cells is needed for its cytotoxicity, so the uptake rate should correlate with the rate of neurodegeneration. Also, the initial step in forming extracellular aggregates cannot occur in the interstitial fluid due to the extremely low concentration of beta-amyloid but can occur intralysosomally. Therefore, the density of extracellular aggregates should positively correlate with the rate of cellular amyloid uptake. The model, which considers that both cytotoxicity and aggregation of beta-amyloid are defined by cellular uptake, successfully reproduces the probability distribution of AD diagnosis in the population. Cellular uptake of beta-amyloid is mediated by endocytosis. Endocytosed beta-amyloid induces lysosomal permeabilization that occurs without plasma membrane damage. Lysosomal permeabilization explains ion disturbances, such as an accumulation of intracellular calcium, caused by cell exposure to extracellular beta-amyloid. Some amyloid fragments, produced from beta-amyloid by lysosomal proteases, can form membrane channels in lysosomal membranes, which are large enough to leak cathepsins to the cytoplasm. Appearance of proteases in the cytoplasm results in necrosis and/or initiation of apoptosis. If the cell survives, the damage of lysosomes leads to autophagy failure and slow recycling of mitochondria, promoting the production of reactive oxygen species and potentiating cell damage.Considering the above, the integrative theory of AD etiology and pathogenesis can be formulated. The etiology of AD is the membrane channel formation by amyloid fragments produced in lysosomes. The pathogenesis includes lysosomal permeabilization by giant membrane channels, which leak lysosomal proteases into the cytoplasm. The correlation between the density of amyloid aggregates and the probability of AD appears because the intensity of cellular uptake defines both aggregation rates in vivo and cytotoxicity of beta-amyloid.The amyloid degradation toxicity hypothesis is the integrative theory of Alzheimer’s disease (AD). It successfully interprets multiple phenomena and paradoxes associated with AD pathobiology at various levels, from molecular and cellular to biomarkers. The hypothesis explains the limitations of currently used biomarkers of AD and proposes etiology-related parameters. These parameters could be measured in humans and become novel diagnostic and prognostic clinical tools. Based on the proposed framework, we foresee the development of effective medications to treat, stall the progression of, or prevent disease development.
... Lysosomal proteases digest beta-amyloid into channel-forming fragments, while lysosomal membrane carries significant negative surface charge. 7 Cellular level Cells exposed to the Aβ42 demonstrate synchronous oscillations of concentrations of several intracellular ions (calcium, sodium, potassium, pH), that matches the concept of an opening of a non-selective membrane channel [18]. ...
... The frequency of channel conductance oscillations that are observed electrophysiologically (millisecond range, [4,5]) does not match the time scale of oscillations in intracellular ion concentrations (seconds to minutes, [18,19]). ...
... The cellular ion concentrations return to their baseline after each individual wave induced by exposure to beta-amyloid [18]. ...
Preprint
Full-text available
Amyloid degradation toxicity hypothesis is the integrative theory of Alzheimer’s disease (AD). It successfully interprets phenomena and paradoxes associated with AD pathobiology. The hypothesis explains the limitations of currently used biomarkers of AD and proposes etiology-related parameters. These parameters could be measured in humans and become novel diagnostic and prognostic clinical tools. Based on the proposed framework, we foresee the development of effective medications to treat, stall the progression of, or prevent the development of the disease. The promise is supported by published preclinical data. This manuscript summarizes the amyloid degradation toxicity hypothesis of Alzheimer’s disease at different levels of organization.
... It is known since the mid-1970s that hormones and neurotransmitters can generate a biphasic cytosolic Ca 2+ signal that reflects a rapid release of Ca 2+ from intracellular stores followed by a Ca 2+ entry from the extracellular fluid. For instance, pioneer works involving 45 Ca flux measurements established that neurotransmitters like acetylcholine can evoke an intracellular Ca 2+ rise accompanied by an entry of Ca 2+ [16]. Thereafter, two seminal findings were reported in the 1980s: (i) the identification of inositol 1,4,5-trisphosphate (IP 3 ) as an intracellular second messenger critical for the release of Ca 2+ from intracellular stores but not for the entry of Ca 2+ [17], and (ii) the identification of intracellular Ca 2+ stores in neurons [18][19][20]. ...
... Therefore, the persistent basal Ca 2+ uptake through constitutive active channels of the cell surface could make SOCC of no use, whatever the cell type, in neurons like in non-neuronal cells. As already pointed out, the question of SOCE in neurons has been the source of some contention [7,45]. However, a growing number of reports has provided experimental evidence for the existence in various neuronal cell types of a Ca 2+ influx triggered in response to the depletion of intracellular Ca 2+ pools and not mediated by VGCCs. ...
Article
Full-text available
The endoplasmic reticulum (ER) is the major intracellular calcium (Ca²⁺) storage compartment in eukaryotic cells. In most instances, the mobilization of Ca²⁺ from this store is followed by a delayed and sustained uptake of Ca²⁺ through Ca²⁺-permeable channels of the cell surface named store-operated Ca²⁺ channels (SOCCs). This gives rise to a store-operated Ca²⁺ entry (SOCE) that has been thoroughly investigated in electrically non-excitable cells where it is the principal regulated Ca²⁺ entry pathway. The existence of this Ca²⁺ route in neurons has long been a matter of debate. However, a growing body of experimental evidence indicates that the recruitment of Ca²⁺ from neuronal ER Ca²⁺ stores generates a SOCE. The present review summarizes the main studies supporting the presence of a depletion-dependent Ca²⁺ entry in neurons. It also addresses the question of the molecular composition of neuronal SOCCs, their expression, pharmacological properties, as well as their physiological relevance.
... Astrocytes are major glial cells in the human brain [3] and their activity is guided by intracellular Ca 2+ signaling [4] which in turn mediates neuronal activity by regulating ambient transmitter and ion recycling [5,6], delivery of energy fuels [7] and the release of transmitter substances [8]. Acute or chronic exposure to soluble and fibril formed Aβ oligomers could raise intracellular calcium levels in several models of AD [9][10][11], and has been found to compromise neuronal survival [12]. Elevated calcium levels owing to Aβ, disrupt the Ca 2+ -dependent release of neurotransmitters such as glutamate, D-serine, and ATP via astrocytes, altering gliotransmission [13]. ...
Article
Full-text available
Astrocytes are the major glial cells in the human brain and provide crucial metabolic and trophic support to neurons. The amyloid-β peptide (Aβ) alter the morphological and functional properties of astrocytes and induce inflammation and calcium dysregulation, contributing to Alzheimer's disease (AD) pathology. Recent studies highlight the role of Toll-like receptor (TLR) 4/nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) signaling in inflammation. Reactive oxygen species (ROS) generated due to Aβ, induce apoptosis in the brain cells worsening AD progression. Astrocytic cell surface receptors, such as purinergic receptors (P2Y1 and P2Y2), metabotropic glutamate receptor (mGLUR)5, α7 nicotinic acetylcholine receptor (α7nAChR), and N-methyl-d-aspartate receptors (NMDARs), have been suggested to interact with inositol trisphosphate receptor (IP3R) on the endoplasmic reticulum (ER) to induce Ca2+ movement from ER to cytoplasm, causing Ca2+ dysregulation. We found that the citrus flavonoid nobiletin (NOB) protected primary astrocytes from Aβ42-induced cytotoxicity and inhibited TLR4/NF-κB signaling in Aβ42-induced primary rat astrocytes. NOB was found to regulate Aβ42-induced ROS levels through Keap1-Nrf2 pathway. The receptors P2Y1, P2Y2, mGLUR5, α7nAChR, and NMDARs induced intracellular Ca2+ levels by activating IP3R and NOB regulated them, thereby regulating intracellular Ca2+ levels. Molecular docking analysis revealed a possible interaction between NOB and IP3R in IP3R regulation. Furthermore, RNA sequencing revealed various NOB-mediated biological signaling pathways, such as the AD-presenilin, AD-amyloid secretase, and Wnt signaling pathway, suggesting possible neuroprotective roles of NOB. To conclude, NOB is a promising therapeutic agent for AD and works by modulating AD pathology at various levels in Aβ42-induced primary rat astrocytes.
... The biological aspects of the influence of proteins involved in neurodegeneration remains a subject of debate. It is known that in AD Aβ inhibits some components of the respiratory chain, thus causing mitochondrial depolarization, and also causes an increase in calcium levels, which leads to oxidative stress [173][174][175]. Disruption of the Tau protein leads to an abnormal distribution of mitochondria [176], a violation of the balance of their fusion and division [177]. ...
Article
Full-text available
Neurodegenerative diseases are currently incurable. Numerous experimental data accu- mulated over the past fifty years have brought us closer to understanding the molecular and cell mechanisms responsible for their development. However, these data are not enough for a complete understanding of the genesis of these diseases, nor to suggest treatment methods. It turns out that many cellular pathologies developing during neurodegeneration coincide from disease to disease. These observations give hope to finding a common intracellular target(s) and to offering a universal method of treatment. In this review, we attempt to analyze data on similar cellular disorders among neurodegenerative diseases in general, and polyglutamine neurodegenerative diseases in particular, focusing on the interaction of various proteins involved in the development of neurodegenerative diseases with various cellular organelles. The main purposes of this review are: (1) to outline the spectrum of common intracellular pathologies and to answer the question of whether it is possible to find potential universal target(s) for therapeutic intervention; (2) to identify specific intracellular pathologies and to speculate about a possible general approach for their treatment.
... Interestingly, a reduction of GSH was observed in the PFC of the 10-month-old 5xFAD mice, while that of the 14-month-old 5xFAD mice was elevated compared to WT. The alteration of GSH levels have been reported in AD brains [34][35][36], suggesting that GSH plays a vital role in the pathological pathway of AD [37]. Although the elevation of GSH in 5xFAD mice remains undetermined, reduction of GSH was consistent with the previous study, which reported a significant decrease of GSH/Cr in the cortical region of the Tg2576 mice compared to the control [12]. ...
Article
Full-text available
This study aimed to investigate morphological and metabolic changes in the brains of 5xFAD mice. Structural magnetic resonance imaging (MRI) and 1H magnetic resonance spectroscopy (MRS) were obtained in 10- and 14-month-old 5xFAD and wild-type (WT) mice, while 31P MRS scans were acquired in 11-month-old mice. Significantly reduced gray matter (GM) was identified by voxel-based morphometry (VBM) in the thalamus, hypothalamus, and periaqueductal gray areas of 5xFAD mice compared to WT mice. Significant reductions in N-acetyl aspartate and elevation of myo-Inositol were revealed by the quantification of MRS in the hippocampus of 5xFAD mice, compared to WT. A significant reduction in NeuN-positive cells and elevation of Iba1- and GFAP-positive cells supported this observation. The reduction in phosphomonoester and elevation of phosphodiester was observed in 11-month-old 5xFAD mice, which might imply a sign of disruption in the membrane synthesis. Commonly reported 1H MRS features were replicated in the hippocampus of 14-month-old 5xFAD mice, and a sign of disruption in the membrane synthesis and elevation of breakdown were revealed in the whole brain of 5xFAD mice by 31P MRS. GM volume reduction was identified in the thalamus, hypothalamus, and periaqueductal gray areas of 5xFAD mice.
... The biological aspects of the influence of proteins involved in neurodegeneration remains a subject of debate. It is known that in AD Aβ inhibits some components of the respiratory chain, thus causing mitochondrial depolarization, and also causes an increase in calcium levels, which leads to oxidative stress [173][174][175]. Disruption of the Tau protein leads to an abnormal distribution of mitochondria [176], a violation of the balance of their fusion and division [177]. ...
Article
Full-text available
Neurodegenerative diseases are currently incurable. Numerous experimental data accumulated over the past fifty years have brought us closer to understanding the molecular and cell mechanisms responsible for their development. However, these data are not enough for a complete understanding of the genesis of these diseases, nor to suggest treatment methods. It turns out that many cellular pathologies eveloping during neurodegeneration coincide from disease to disease. These observations give hope to finding a common intracellular target(s) and to offering a universal method of treatment. In this review, we attempt to analyze data on similar cellular disorders among neurodegenerative diseases in general, and polyglutamine neurodegenerative diseases in particular, focusing on the interaction of various proteins involved in the development of neurodegenerative diseases with various cellular organelles. The main purposes of this review are: (1) to outline thes pectrum of common intracellular pathologies and to answer the question of whether it is possible to find potential universal target(s) for therapeutic intervention; (2) to identify specific intracellular pathologies and to speculate about a possible general approach for their treatment.
Article
Full-text available
Neurodegenerative diseases are chronic conditions occurring when neurons die in specific brain regions that lead to loss of movement or cognitive functions. Despite the progress in understanding the mechanisms of this pathology, currently no cure exists to treat these types of diseases: for some of them the only help is alleviating the associated symptoms. Mitochondrial dysfunction has been shown to be involved in the pathogenesis of most the neurodegenerative disorders. The fast and transient permeability of mitochondria (the mitochondrial permeability transition, mPT) has been shown to be an initial step in the mechanism of apoptotic and necrotic cell death, which acts as a regulator of tissue regeneration for postmitotic neurons as it leads to the irreparable loss of cells and cell function. In this study, we review the role of the mitochondrial permeability transition in neuronal death in major neurodegenerative diseases, covering the inductors of mPTP opening in neurons, including the major ones—free radicals and calcium—and we discuss perspectives and difficulties in the development of a neuroprotective strategy based on the inhibition of mPTP in neurodegenerative disorders.
Article
Full-text available
In Alzheimer's disease (AD), abnormal accumulations of beta-amyloid are present in the brain and degenerating neurons exhibit cytoskeletal aberrations (neurofibrillary tangles). Roles for beta-amyloid in the neuronal degeneration of AD have been suggested based on recent data obtained in rodent studies demonstrating neurotoxic actions of beta- amyloid. However, the cellular mechanism of action of beta-amyloid is unknown, and there is no direct information concerning the biological activity of beta-amyloid in human neurons. We now report on experiments in human cerebral cortical cell cultures that tested the hypothesis that beta-amyloid can destabilize neuronal calcium regulation and render neurons more vulnerable to environmental stimuli that elevate intracellular calcium levels. Synthetic beta-amyloid peptides (beta APs) corresponding to amino acids 1–38 or 25–35 of the beta-amyloid protein enhanced glutamate neurotoxicity in cortical cultures, while a peptide with a scrambled sequence was without effect. beta APs alone had no effect on neuronal survival during a 4 d exposure period. beta APs enhanced both kainate and NMDA neurotoxicity, indicating that the effect was not specific for a particular subtype of glutamate receptor. The effects of beta APs on excitatory amino acid (EAA)-induced neuronal degeneration were concentration dependent and required prolonged (days) exposures. The beta APs also rendered neurons more vulnerable to calcium ionophore neurotoxicity, indicating that beta APs compromised the ability of the neurons to reduce intracellular calcium levels to normal limits. Direct measurements of intracellular calcium levels demonstrated that beta APs elevated rest levels of calcium and enhanced calcium responses to EAAs and calcium ionophore. The neurotoxicity caused by EAAs and potentiated by beta APs was dependent upon calcium influx since it did not occur in calcium-deficient culture medium. Finally, the beta APs made neurons more vulnerable to neurofibrillary tangle-like antigenic changes induced by EAAs or calcium ionophore (i.e., increased staining with tau and ubiquitin antibodies). Taken together, these data suggest that beta-amyloid destabilizes neuronal calcium homeostasis and thereby renders neurons more vulnerable to environmental insults.
Article
Full-text available
A new family of highly fluorescent indicators has been synthesized for biochemical studies of the physiological role of cytosolic free Ca2+. The compounds combine an 8-coordinate tetracarboxylate chelating site with stilbene chromophores. Incorporation of the ethylenic linkage of the stilbene into a heterocyclic ring enhances the quantum efficiency and photochemical stability of the fluorophore. Compared to their widely used predecessor, “quin2”, the new dyes offer up to 30-fold brighter fluorescence, major changes in wavelength not just intensity upon Ca2+ binding, slightly lower affinities for Ca2+, slightly longer wavelengths of excitation, and considerably improved selectivity for Ca2+ over other divalent cations. These properties, particularly the wavelength sensitivity to Ca2+, should make these dyes the preferred fluorescent indicators for many intracellular applications, especially in single cells, adherent cell layers, or bulk tissues.
Article
Full-text available
Aβ binds Zn2+, Cu2+, and Fe3+ in vitro, and these metals are markedly elevated in the neocortex and especially enriched in amyloid plaque deposits of individuals with Alzheimer's disease (AD). Zn2+ precipitates Aβ in vitro, and Cu2+ interaction with Aβ promotes its neurotoxicity, correlating with metal reduction and the cell-free generation of H2O2 (Aβ1–42 > Aβ1–40 > ratAβ1–40). Because Zn2+ is redox-inert, we studied the possibility that it may play an inhibitory role in H2O2-mediated Aβ toxicity. In competition to the cytotoxic potentiation caused by coincubation with Cu2+, Zn2+ rescued primary cortical and human embryonic kidney 293 cells that were exposed to Aβ1–42, correlating with the effect of Zn2+ in suppressing Cu2+-dependent H2O2formation from Aβ1–42. Since plaques contain exceptionally high concentrations of Zn2+, we examined the relationship between oxidation (8-OH guanosine) levels in AD-affected tissue and histological amyloid burden and found a significant negative correlation. These data suggest a protective role for Zn2+in AD, where plaques form as the result of a more robust Zn2+ antioxidant response to the underlying oxidative attack.
Article
Given the increasing evidence of oxidative stress in AD brain3,4 and studies from different perspectives that appear to show a converging, central role for Aβ in the pathogenesis and etiology of AD,21,22,28 insight into Aβ-associated free radical oxidative stress will likely lead to a greater understanding of AD and, potentially, to better therapeutic strategies in this disorder. This article outlined methods to investigate markers of oxidative stress induced by Aβ in brain membrane systems.3 Especially important are markers for protein oxidation, lipid peroxidation, and ROS generation by Aβ. Oxidative stress and its sequelae are likely related to both necrotic and apoptotic mechanisms of neurotoxicity,4 and Aβ-associated free radical oxidative stress may be of fundamental importance in Alzheimer's disease etiology and pathogenesis.3,4 The methods described here provide some means for investigating this possibility.
Article
Effective drugs are not available to protect against β-amyloid peptide (Aβ)-induced neurotoxicity. Cortical neurons from rat embryos were treated with the toxic fragment Aβ25-35 at 1 µM in the presence or absence of flupirtine, a triaminopyridine, successfully applied clinically as a nonopiate analgesic drug. Five days later 1 µM Aβ25-35 caused reduction of cell viability to 31.1%. Preincubation of cells with flupirtine (1 or 5 µg/ml) resulted in a significant increase of the percentage of viable cells (74.6 and 65.4%, respectively). During incubation with Aβ25-35 the neurons undergo apoptosis as determined by appearance of the characteristic stepladder-like DNA fragmentation pattern and by the TUNEL technique. Aβ25-35-induced DNA fragmentation could be abolished by preincubation of the cells with 1 µg/ml flupirtine. Incubation with Aβ25-35 reduces the intraneuronal level of GSH from 21.4 to 7.4 nmol/106 cells. This depletion could be partially prevented by preincubation of the cells with flupirtine. Thus, flupirtine may be adequate for the treatment of the neuronal loss in Alzheimer's disease (where Aβ accumulates in senile plaques) and probably other neurological diseases such as amyotrophic lateral sclerosis.
Article
Perturbations to glutathione (GSH) metabolism may play an important role in neurodegenerative disorders such as Alzheimer's, Parkinson's, and prion diseases. A primary function of GSH is to prevent the toxic interaction between free radicals and reactive transition metals such as copper (Cu). Due to the potential role of Cu in neurodegeneration, we examined the effect of GSH depletion on Cu toxicity in murine primary neuronal cultures. Depletion of cellular GSH with L-buthionine-[S,R]-sulfoximine resulted in a dramatic potentiation of Cu toxicity in neurons without effect on iron (Fe) toxicity. Similarly, inhibition of glutathione reductase (GR) activity with 1,3-bis(2-chloroethyl)-1-nitrosurea also increased Cu toxicity in neurons. To determine if the Alzheimer's amyloid-β (Aβ) peptide can affect neuronal resistance to transition metal toxicity, we exposed cultures to nontoxic concentrations of Aβ25-35 in the presence or absence of Cu or Fe. Aβ25-35 pretreatment was found to deplete neuronal GSH and increase GR activity, confirming the ability of Aβ to perturb neuronal GSH homeostasis. Aβ25-35 pretreatment potently increased Cu toxicity but had no effect on Fe toxicity. These studies demonstrate an important role for neuronal GSH homeostasis in selective protection against Cu toxicity, a finding with widespread implications for neurodegenerative disorders.
Article
Amyloid β protein (Aβ), the central constituent of senile plaques in Alzheimer's disease (AD) brain, is known to exert toxic effects on cultured neurons. The role of the voltage-sensitive Ca2+ channel (VSCC) in β(25–35) neurotoxicity was examined using rat cultured cortical and hippocampal neurons. When L-type VSCCs were blocked by application of nimodipine, β(25–35) neurotoxicity was attenuated, whereas application of ω-conotoxin GVIA (ω-CgTX-GVIA) or ω-agatoxin IVA (ω-Aga-IVA), the blocker for N- or P/Q-type VSCCs, had no effects. Whole-cell patch-clamp studies indicated that the Ca2+ current density of β(25–35)-treated neurons is about twofold higher than that of control neurons. Also, β(25–35) increased Ca2+ uptake, which was sensitive to nimodipine. The 2′,7′-dichlorofluorescin diacetate assay showed the ability of β(25–35) to produce reactive oxygen species. Nimodipine had no effect on the level of free radicals. In contrast, vitamin E, a radical scavenger, reduced the level of free radicals, neurotoxicity, and Ca2+ uptake. These results suggest that β(25–35) generates free radicals, which in turn, increase Ca2+ influx via the L-type VSCC, thereby inducing neurotoxicity.